1 Introduction

With the dramatic developments in observational technologies, a large number of new phenomena have been discovered in various astronomical experiments in the past decade. Our understanding of the very nature of the Universe, especially of the exotic astrophysical objects including black holes, supernovae, neutrons, blazars, dark matter (DM), and active galactic nuclei, has been greatly advanced. Specifically, high-energy gamma-ray observations provide the only available probe to identify cosmic-ray sources which could reveal unique information about the exotic celestial objects, as gamma rays travel in a straight line from the source without deflection by a galactic magnetic field [1,2,3,4]. However, many cosmic-ray sources have physical origins that are still under discussion [5, 6].

Also, accumulated gravitational wave (GW) events detected by the LIGO/Virgo Collaboration ushered in a new era in observational astronomy that shed light on the formation of exotic astrophysical objects [7], and have been used to test general relativity in the unexplored strong-gravity regime [8]. Hence, the development of multi-messenger observation, which is a joint observation of cosmic rays, neutrinos, photons, and GWs, could provide us unique insights into the properties of astrophysical sources and source populations in our Universe [9]. With multi-messenger observation, there is a growing interest in searching for exotic celestial bodies, such as quark stars [10], boson stars [11], dark stars [12], and antistars [13]. Such a search can reveal new aspects of fundamental science and serve as new probes into the primordial Universe.

In this paper, we propose a novel class of exotic celestial objects filled by primordial black holes (PBHs). These objects appear as bubbles of stellar size and can emit gamma rays and GWs through the processes of Hawking radiation and binary mergers, respectively. Theoretically, these stellar bubbles can be generated from some new-physics phenomena that might have occurred in the primordial Universe, such as quantum tunnelings during or after inflation [14, 15], multi-stream inflation [16], and inhomogeneous baryogenesis [17]. In these cases, before the bubble-wall tension vanishes, the field values differ between the inside and outside of the bubble. Such difference can result in different local physics inside the bubble (for the PBH case, see [18, 19] for details), namely the production rate of the exotic species of matter illustrated below, which indicates the existence of the “island universes,” as a baby version of a multiverse. If the size of these bubbles is small enough, say, smaller than the resolution of current telescopes, they behave as exotic celestial objects.

Observational consequences of exotic stellar bubbles can arise from (i) decay, annihilation, or interaction of exotic matter inside the bubble, for example, unstable particles, textures, monopoles, cosmic strings, domain walls, and PBHs, that can yield cosmic rays at high energy scales and hence may address the puzzle of ultrahigh-energy cosmic rays, such as PeV gamma rays discovered in the recent Large High Altitude Air Shower Observatory (LHAASO) observation [20]; (ii) interaction at the interface of the bubble, for example, if the stellar bubble is made of antimatter, where the effects can be analogous to [13]. If there are overdense regions inside the bubble, a certain amount of PBHs can be formed due to the local gravitational collapse [21,22,23], resulting in a wide range of PBH masses that are quite different from the astrophysical black holes [24,25,26]. When the masses of these PBHs are small enough, their Hawking radiation becomes sufficient to generate observable electromagnetic (EM) signals. Additionally, PBHs can also cluster to form binaries and generate GWs from their mergers [27, 28]. This opens the door to searching for the nontrivial observational signals from these exotic celestial bodies, as we will investigate in this paper.

2 EM radiation

In [29, 30], Hawking remarkably found that black holes emit radiation in the black-body form. Theoretically, PBHs satisfy a certain mass distribution depending on the underlying generation mechanisms [31,32,33,34], and hence one can derive the EM radiation contributed by all possible mass scales. Let us focus our interest on neutral PBHs of the Schwarzschild type. The time-dependent physical number density \(n_i\) of elementary particle i emitted by a distribution of PBHs per unit time and per unit energy can be determined by

$$\begin{aligned} {\mathrm {d}^2 n_i \over \mathrm {d}t \mathrm {d}E}(E) = \int _{M_\text {min}}^{M_\text {max}} \frac{\mathrm {d}^2 N_i}{\mathrm {d}t \mathrm {d}E}(E,M) n_\text {PBH}(M) \mathrm {d}M , \end{aligned}$$
(1)

where \(\mathrm {d}^2 N_i / \mathrm {d}t \mathrm {d}E\) is the Hawking instantaneous emission rate for particle i. The lower mass limit \(M_\text {min}\) is usually set to the Planck mass \(M_\text {pl} \simeq 2 \times 10^{-5}\) g, while the upper one \(M_\text {max}\) can be infinity.

In general, the mass distribution of PBHs can be described by the differential physical mass function \(n_\text {PBH}(M) \equiv \mathrm {d} n_\text {PBH} / \mathrm {d} M\), where \(\mathrm {d} n_\text {PBH}\) is the physical number density of PBHs in the mass range \((M, M + \mathrm {d}M)\). For heavy PBHs which survive at present (the formation mass \(M_f \gtrsim 10^{15}\) g), the current energy fraction \(\tilde{f}_{\text {PBH}}\) of PBHs over the DM component inside PBH bubbles is given by the integral \(\tilde{f}_{\text {PBH}} \equiv \tilde{\Omega }_\text {PBH} / \tilde{\Omega }_\text {DM} = \int _{M_\text {min}}^{M_\text {max}} \psi (M) \mathrm {d}M\), where \(\tilde{\Omega }_\text {PBH}\) and \(\tilde{\Omega }_\text {DM}\) are the current normalized density of PBHs and DM inside PBH bubbles, respectively. Note that we use the notations \(\Omega _{\text {rad}}\) and \(\Omega _{\text {DM}}\) to denote the corresponding quantities defined over the whole observable Universe. The current mass function \(\psi (M)\) is defined as \(\psi (M) \equiv M n_\text {PBH}(M) / \rho _\text {DM}\), where the scale factor at present takes \(a(t_0) = 1\). If the PBHs were from inflationary fluctuations, they naturally satisfy a lognormal mass function [26, 35],

$$\begin{aligned} \psi _\text {LN}(M) = {\tilde{f}_{\text {PBH}} \over \sqrt{2 \pi } \sigma M} \exp \Bigg [ - { \ln ^2(M/M_\text {pk}) \over 2 \sigma ^2 } \Bigg ] , \end{aligned}$$
(2)

where the subscript “LN” denotes the lognormal type. Given \(\tilde{f}_{\text {PBH}}\), there are two parameters in \(\psi _\text {LN}(M)\): \(\sigma \) describes the width of mass distribution, and \(M_\text {pk}\) is the mass at which the function \(M \psi _\text {LN}(M)\) (the DM faction in PBHs at the logarithmic interval around M) peaks. Since light PBHs (\(M_f \lesssim 10^{15}\) g) would have evaporated at an earlier time, the present mass function could be deformed from the initial shape, especially in the low-mass tail [36].

Let us emphasize that due to the clustering of PBHs inside PBH bubbles, the local density \(\tilde{f}_{\text {PBH}}\) can be amplified. Taking multi-stream inflation as an example, we use \(\beta _1\) to denote the volume fraction of PBH bubbles in the observed Universe. If we assume that the dark matter energy density is the same inside PBH bubbles and the whole observable Universe, i.e., \(\Omega _{\text {DM}} = \tilde{\Omega }_{\text {DM}}\), we have the following relation: \(\tilde{f}_{\text {PBH}} = \beta _1^{-1} f_{\text {PBH}}\), where \(f_{\text {PBH}} \equiv \Omega _{\text {rad}}/\Omega _{\text {DM}}\) is the standard definition. In the multi-stream inflation, the probability \(\beta _1 \ll 1\), so that \(\tilde{f}_{\text {PBH}}\) is largely amplified. According to the current constraints on \(f_{\text {PBH}}\) [37], \(f_{\text {PBH}} \sim \mathcal {O}(10^{-3})\) for the wide mass window \(10^{-16}\)\(10^{1} M_\odot \), it is straightforward to know that \(\tilde{f}_{\text {PBH}} > 1\) requires \(\beta _1 < 10^{-3}\), corresponding to the size of PBH bubbles: \(V_\text {bubb} < 3\) Mpc, which is quite easily satisfied in the multi-stream inflation. Actually, for a stellar bubble, the amplification could be much larger than \(10^3\). The number of PBHs in the clustering can be estimated by

$$\begin{aligned} N = \left( \frac{k_2}{k_1} \right) ^3 \tilde{\beta }. \end{aligned}$$
(3)

Here, \(1/k_1\) denotes the comoving scale which exits the horizon at the time of bifurcation in the multi-stream inflation, and \(1/k_2\) denotes the comoving scale which exits the horizon at the time of PBH formation during inflation. \(\tilde{\beta }\) is the initial PBH abundance inside a PBH bubble that corresponds to the current local density \(\tilde{f}_{\text {PBH}}\). The scale \(1/k_1\) determines the size of the exotic stellar bubble, which depends on the potential of the multi-stream inflation. In practice, we take \(k_1^{-1} = 1 \, \mathrm {Mpc}\) as an example. \(k_2\) is related to the mass of PBHs as [38]

$$\begin{aligned} k_2 \simeq 7.5 \times 10^5 \, \mathrm {Mpc}^{-1}\left( \frac{M_\mathrm {PBH}}{30 M_{\odot }} \right) ^{-1/2} , \end{aligned}$$
(4)

where \(M_{\odot }\) is the solar mass. We take \(M_\mathrm {PBH} = 10^{15} \mathrm {g}\), which gives \(k_2 \sim \mathcal {O}(10^{15}) \mathrm {Mpc}^{-1}\), and the number of PBHs in the clustering is estimated as

$$\begin{aligned} N \simeq 2 \times 10^{47} \tilde{\beta }. \end{aligned}$$
(5)

The calculation of PBH abundance in PBH stellar bubbles is similar to the standard calculation [26]. For example, for the critical collapse model [37, 39, 40], the local abundance is given by \(\tilde{\beta } \simeq k \sigma ^{2 \gamma } \text {erfc} \left( \frac{\delta _c}{\sqrt{2} \sigma }\right) \). Hence, \(N > 10^{17}\) if \(\tilde{\beta } > 10^{-30}\), which is easily realized in various PBH formation mechanisms [37]. Thus, the number N of PBHs inside a PBH bubble in general is largely amplified, and it is safe to regard N as a free parameter in the following discussions.

The extragalactic gamma-ray background (EGB) is an important constraint for evaporating PBHs [41]. With the advent of gamma-ray experiments in various energy ranges, such as the Imaging Compton Telescope (COMPTEL) for 0.8–30 MeV [42], the Energetic Gamma Ray Experiment Telescope (EGRET) for 20 MeV \(\sim 30\) GeV [43], and the Fermi Large Area Telescope (Fermi LAT) for 20 MeV \(\sim 300\) GeV [44], PBHs within the mass range \(10^{14} \sim 10^{16}\) g are severely constrained and cannot provide the dominant contribution to DM [45,46,47]. Moreover, the galactic gamma-ray background [36] and gamma-ray bursts [48, 49] can be applied to constrain PBHs.

Now, we study the EM signals from a single PBH stellar bubble and the detectability in gamma-ray channels. Consider an initial lognormal distributed PBH stellar bubble located at redshift z. The key EM observable is the photon flux detected on Earth \(F(\tilde{E},t) \equiv \tilde{E}^2 \mathrm {d}^2\tilde{n}_\gamma / \mathrm {d} \tilde{E} \mathrm {d}t\), where \(\tilde{E}\) and \(\tilde{n}_\gamma \) are the photon energy and number density observed on Earth. This observed photon flux is related to the intrinsic luminosity L(Ez) of Hawking radiation from the PBH bubble located at z via

$$\begin{aligned} F(E,z) = { L(E(1 + z),z) \over 4 \pi d_L^2(z) } , \end{aligned}$$
(6)

where L(Ez) is determined by the emitted photon physical number density per unit energy and per unit time \(\mathrm {d}^2 n_\gamma / \mathrm {d}t \mathrm {d}E\). It can be numerically computed as shown in Fig. 1 for the peak masses \(M_\text {pk} = 10^{13}, 10^{15}, 10^{17}\) g and the energies of emitted photons \(E = 10, 100\) GeV, respectively. The flat parts in this panel are caused by small masses radiating at the early stage of Hawking radiation. The luminosity distance in Eq. (6) is given by \(d_L(z) = (1 + z) \int _0^z \mathrm {d}\tilde{z} / H(\tilde{z})\), which accounts for the redshift of the photon energy and apparent emission rate [50, 51]. Combining (1) and (6) yields the observed photon fluxes in various photon energy ranges from a single PBH bubble located at z, as reported in Fig. 2. It shows that for the small peak mass \(M_\text {pk} = 10^{13}\) g, the observable fluxes of the PBH bubble of redshift \(z\gtrsim 1\) would increase to some extent in contrast to the larger peak mass \(M_\text {pk} = 10^{15}, 10^{17}\) g, due to the evaporation of a majority of PBH bubbles at high redshifts.

Fig. 1
figure 1

The spectrum \(\mathrm {d}^2 n_\gamma / \mathrm {d}t \mathrm {d}E\) of a PBH bubble as a function of cosmic time t for peak masses \(M_\text {pk} = 10^{13}, 10^{15}, 10^{17}\) g and photon energies \(E = 10, 10amma-ray si0\) GeV, respectively. For simplicity, we have set \(\tilde{f}_{\text {PBH}} =1\) and \(\sigma =1\)

Fig. 2
figure 2

The observed photon flux F of a PBH bubble of physical volume 1 Mpc\(^3\) located at different redshifts for various peak masses and photon energies

Fig. 3
figure 3

The parameter space of the peak mass \(M_\text {pk}\) and bubble mass \(M_\text {bub}\) of a PBH bubble allowed by Fermi LAT (the shaded regions) for redshifts \(z=0.1, 1, 10\) and photon energies \(E= 10, 100\) GeV, respectively

Fig. 4
figure 4

The parameter space of the redshift z and bubble mass \(M_\text {bub}\) allowed by Fermi LAT (the shaded regions) with the photon energy \(E = 10\) GeV and peak masses \(M_\text {pk} = 10^{13}, 10^{15}, 10^{17}\) g, respectively

3 Gamma-ray signals

Recall that the PBH mass function in (2) relies on three parameters, i.e., \(\tilde{f}_{\text {PBH}}\), \(\sigma \), and \(M_\text {pk}\). In the case where PBHs cluster as a single bubble, we generalize the concept of \(\tilde{f}_{\text {PBH}}\) to be the energy density of the PBHs inside the bubble, which is convenient for our calculation. Furthermore, the intrinsic luminosity of Hawking radiation involves two other parameters of a bubble, i.e., its physical volume V and redshift z. For observable interest, we combine \(\tilde{f}_{\text {PBH}}\) and V to yield the initial total mass of a PBH bubble \(M_\text {bub}\). Thus, we get four parameters: \(M_\text {bub}\), z, \(M_\text {pk}\), and \(\sigma \). Using the point-source differential sensitivity in the 10-year observation of Fermi LAT for a high Galactic latitude (around the North Celestial Pole) source [44], we numerically derive the experimentally allowed parameter spaces for (\(M_\text {bub}\), \(M_\text {pk}\)) and (\(M_\text {bub}\), z) by the shaded regions in Figs. 3 and 4, respectively. Setting \(\sigma = 1\), for the given redshifts and photon energies, the lower bound of the parameter space for \(M_\text {pk}\) and \(M_\text {bub}\) is around \(M_\text {pk} \simeq 10^{15}\) g, which is the mass scale evaporating at present, and the corresponding bubble mass is \(M_\text {bub} \simeq 10^{32} \text {g}\). For the smaller or larger values of \(M_\text {pk}\), the corresponding Hawking radiation has either decayed out already or not yet become efficient, and thus the bubble mass \(M_\text {bub}\) needs to be heavy enough to yield observable evidence. Additionally, Figs. 3 and 4 indicate that the closer the stellar PBH bubbles are to the Earth, the easier they can be probed. The “plateau” for the case of \(M_\text {pk} = 10^{13}\) g and \(z \gtrsim 1\) that appears in Fig. 4 occurs for the same reason as the fluxes of \(M_\text {pk} = 10^{13}\) g shown in the upper right panel.

The LHAASO experiment recently reported the astonishing detection of 12 ultrahigh-energy gamma-ray sources [20], which indicates some high-energy physics in stellar objects. While the astrophysical sources responsible for these events are under debate [1, 2, 52], the possibility of their origin as exotic celestial objects including the PBH stellar bubble warrants further examination. Therefore, we confront our scenario with the latest LHAASO data and report the numerical results in Fig. 5. For connecting the gamma-ray spectrum and the present PBH mass distribution, the present lognormal distribution is considered instead of a primordial one. Accordingly, \(\tilde{M}_\text {pk}\) denotes the peak mass of the present lognormal distribution, and \(\tilde{M}_{\mathrm {bub}}\) denotes the present stellar bubble mass. In the numerical calculations, we set \(\sigma = 1\). Varying \(\sigma \) may lead to better fits to observations, which we leave to future work. Our results illustrate well that the PBH stellar bubbles can provide a decent fit, as explicitly shown in the plot.

Fig. 5
figure 5

The fit of three different PBH stellar bubbles to the data by LHAASO. The observation distance of ultrahigh-energy gamma-ray sources \(\mathrm {J}2226+6057\), \(\mathrm {J}1908+0621\), and \(\mathrm {J}1825-1326\) are 0.8, 2.37, and 1.55 \(\mathrm {kpc}\) [53, 54], respectively. The blue curves are numerical results produced using the publicly available code BlackHawk [55], and the black curves are the superpositions of analytic black body spectra normalized by the BlackHawk calculation

4 GW signals

The evolution of PBH binaries produces GWs. For those PBHs whose masses are less than \(10^4 M_{\odot }\), the associated GWs become very weak for observations. Accordingly, the main observational channel of GWs is to follow the evolution of PBH binaries when considering heavy masses. A PBH binary in a stellar bubble forms from two nearby black holes and decouples from the Hubble flow. Given the initial separation x in a binary with total mass \(M = m_1 + m_2\), the binary system forms at \(z \approx 3(1 + z_{\mathrm {eq}})/\lambda - 1\), where \(z_{\mathrm {eq}}\approx 3000\) is the redshift at matter-radiation equality in the \(\Lambda \mathrm {CDM}\) cosmology with \(\Omega _{\mathrm {M}} = 0.315\), \(\Omega _{\Lambda } = 0.685\) [60], and \(\lambda \equiv 8 \pi \rho _{\mathrm {eq}} x^3/3 M\) [61].

The GW energy spectrum from a single PBH stellar bubble is

$$\begin{aligned} \Omega _{\mathrm {GW}}(f) = \frac{1}{\rho _c}\frac{1}{4 \pi d_L^2} f_r \frac{dE_{\mathrm {GW}}}{df_r} R , \end{aligned}$$
(7)

where f and \(f_r\) are the GW frequency in the observed frame and binaries rest frame, respectively, with \(f = f_r/(1 + z)\). \(\rho _{c} = 3 H_0^2/8 \pi G\) is the critical density of the Universe, and R is the comoving merger rate of PBH binaries [62]. \(dE_{\mathrm {GW}}/df_r\) is the energy emission per frequency interval, which is parameterized by [63,64,65]

$$\begin{aligned}&\frac{dE_{\mathrm {GW}}}{df_r}= (\pi G)^{2/3}\mathcal {M}_c^{5/3}\nonumber \\&\quad \times \left\{ \begin{aligned}&f_r^{-1/3} \qquad ,~f_r< f_1 \\&f_r^{2/3} f_1^{-1} \qquad ,~f_1 \le f_r<f_2 \\&f_4^4 f_r^2\big [ f_1 f_2^{4/3} \big ( 4(f_r \!-\! f_2)^2 \!+\! f_4^2 \big )^2 \big ]^{-1},~f_2 \!\le \! f_r \!<\! f_3 \end{aligned} \right. \nonumber \\ \end{aligned}$$
(8)

where \(\mathcal {M}_c \equiv (m_1 m_2)^{3/5}/M^{1/5}\) is the chirp mass of the binary system and \(f_i = (a_i \eta ^2 + b_i \eta +c_i)/\pi G M\). The symmetric mass ratio \(\eta \) is defined as \(\eta \equiv m_1 m_2/M^2\), and the coefficients \(a_i,~b_i,~c_i\) can be found in Table 1 of [66]. In Eq. (8), the emitted GW energy is proportional to \(\mathcal {M}_c^{5/3}\), and thus the contributions of light PBH binaries to GWs are negligible.

Combining (7) and (8) yields the GW spectrum from a single PBH stellar bubble and detectable mass parameters region in Figs. 6 and 7, respectively. In Fig. 6, with the same \(M_\mathrm {bub}\), the larger \(M_\mathrm {pk}\), the more GWs that are emitted at the same frequency. Bubbles with peak mass around \(10^{15} \mathrm {g}\) emit weak GWs that are difficult to probe within \((10^{-9}, 10^{2}) \mathrm {Hz}\), which is also shown in Fig. 7, where the \(M_\mathrm {bub}\) should be larger than \(10^{53} \mathrm {g}\) with \(M_\mathrm {pk}\) less than \(10^{23} \mathrm {g}\) at \(0.1 \mathrm {Hz}\) for \(\tilde{f}_{\mathrm {PBH}} = 10^3\) and \(z = 0.01\).

Fig. 6
figure 6

The GW spectrum of a PBH stellar bubble. We set \(\tilde{f}_{\text {PBH}} = 1\) in the bubble and \(z = 0.01\). The mass distribution of PBHs is lognormal with \(\sigma = 1\) and \(M_\text {pk} = 10^{34}, 10^{36}, 10^{38}\mathrm {g}\) for red, orange, and blue shadow regions, respectively. The lower and upper solid curves denote the total PBH mass in a bubble, which are set as \(10^{45} \mathrm {g}\) and \(10^{48} \mathrm {g}\), respectively. Sensitivity curves of SKA [56], LISA [57], BBO [58], and LIGO [59] are plotted

Fig. 7
figure 7

The parameter space of the bubble mass and peak mass of a PBH bubble that can be probed by BBO at \(0.1 \mathrm {Hz}\) with \(\Omega _{\mathrm {GW}}h^2 = 2.9 \times 10^{-15}\). The solid and dashed lines denote \(\tilde{f}_{\text {PBH}} = 10, 10^3\), respectively. Blue and red shaded regions denote \(z = 0.001\), 0.01, respectively

While PBH stellar bubbles with strong Hawking radiation can hardly be probed by the current GW surveys, they may be detectable in some ultrahigh-frequency GW experiments. Note that the peak energy density in the GW spectrum with different peak mass is similar. In Eq. (8), the peak energy radiation occurs at \(f_r = f_2\). The comoving merger rate of binaries scales as \(R \sim M^{-32/37}\); see [19, 28, 62] for details. Thus, the total radiation power in binaries’ rest frame is \(R f_r dE_{\mathrm {GW}}/df_r \sim M^{5/37} \eta (a_2 \eta ^2 + b_2 \eta +c_2)^{5/3}/(a_1 \eta ^2 + b_1 \eta +c_1)\), which depends on symmetric mass ratio \(\eta \) and weakly depends on M. Due to the \(M_\text {pk}\) independence in \(\eta \) distribution, the peak energy density in the GW spectrum is weakly dependent on the peak mass in PBH mass distribution. This provides the possibility for detecting GW signals of a strong-Hawking-radiation PBH stellar bubble in the ultrahigh-frequency range.

Apart from the GW signals from the PBH binaries inside PBH stellar bubbles as we have discussed above, the dynamics of bubbles can also produce the stochastic GW background, e.g., the bubble nucleation, through the first-order phase transition, bubble expansion, and the collision of bubbles [67]. We assume that PBH bubbles are a set of isolated spherical bubbles in this work, which cannot produce GWs through nucleation and expansion processes. The dominant stochastic GW background from the bubble dynamics channel is therefore from the collision of bubbles. As for the phase transition during the inflation, the scalar field plays an essential role in producing the GW background during the collision of bubbles [68, 69]. The resulting GW energy spectrum \(\Omega _{\mathrm {GW}}\) follows \(\Omega _{\mathrm {GW}}(f) \propto f^q\), where the spectral indices \(q = 2.8, -1\) in the limits of low and high frequencies, respectively [70], while for the phase transition occurring after the inflation, the GW signals are also contributed by the sound waves [71] and magnetohydrodynamic (MHD) turbulence [72] in the plasma. The GW energy spectral indices from sound waves in the plasma take \(q = 3, -3\) for the low and high frequencies, respectively [73], while the MHD turbulence produces GWs with spectral indices \(q = 3, -5/3\) in the low and high frequencies, respectively [74]. Hence, the GW signals produced from the bubble dynamics can be distinguished from the signals from the PBH binary mergers inside the PBH bubbles by using the future GW observations (see Refs. [75, 76] and references therein). The potential effects of the dynamics of PBH bubbles on GW signals require further investigation in follow-up work.

5 Concluding remarks

To conclude, we propose the hypothetical possibility of stellar bubbles, which are star-like objects in the sky with exotic features. We focus on a specific class of exotic stellar bubbles filled by lognormally distributed PBHs and analyze their signatures through both the EM and GW observational windows. For the EM channel, a bubble dominated by light PBHs can yield detectable gamma-ray spectra via Hawking radiation. The peak mass \(M_\text {pk} \sim 10^{15}\) g and bubble mass \(M_{\mathrm {bub}} \sim 10^{32}\) g can be related to the 10–100 GeV detection band of Fermi LAT (Figs. 3, 4). Impressively, this scenario can achieve a decent fit to the ultrahigh-energy gamma-ray events discovered by LHAASO and hence hints at the existence of PBH stellar bubbles with the present peak mass \(\tilde{M}_\text {pk} \sim 10^{10} \mathrm {g}\) (see Fig. 5). For the GWs channel, we find that massive PBH binaries with \(M_\text {pk} \sim 10^{34}\)\(10^{38}\) g and \(M_{\mathrm {bub}} \sim 10^{45}\)\(10^{48}\) g can produce detectable GWs within in the frequency band of LISA and BBO (see Fig. 6).

We note that EM and GW signals are complementary for light and heavy bubbles. The search for light PBH bubbles may find promise by observing the amount of gamma-ray sources whose spectra follow Hawking spectra. For heavy bubbles, they can be more accessible by GW astronomy. These combined limits can further constrain the mass function of the PBHs and thus infer the PBH formation in the Universe. Accordingly, if such PBH stellar bubbles were observed, this would serve as a novel window to probe the very early Universe.

Additionally, cosmic neutrinos and ultrahigh-frequency GWs could also be generated from these exotic stellar bubbles. The IceCube Neutrino Observatory has confirmed high-energy cosmic neutrinos as key messengers where wavelengths are opaque to EM signals [77]. On the other hand, the PBH bubbles may leave significant GWs of extremely high frequency, which compels the development of ultrahigh-frequency GW technology [78]. We end by mentioning that, instead of PBH stellar bubbles, there exist other types of bubbles that produce observable signals, such as collision of cosmic strings [79] and domain walls [80] in the bubbles, and matter–antimatter annihilation on boundaries [81]. These exotic stellar bubbles from the primordial Universe can unveil rich physics hidden in the light of stars, which deserve to be explored in the future.